26 Phys Sta Solid b 244 4106 4111 2007


phys. stat. sol. (b) 244, No. 11, 4106 4111 (2007) / DOI 10.1002/pssb.200776208
Electronic properties of graphene
K. S. Novoselov*, 1, S. V. Morozov2, T. M. G. Mohinddin3, L. A. Ponomarenko1,
D. C. Elias1, R. Yang3, I. I. Barbolina4, P. Blake5, T. J. Booth5, D. Jiang5, J. Giesbers6,
E. W. Hill7, and A. K. Geim1
1
School of Physics and Astronomy, University of Manchester, Oxford Road, Manchester, M13 9PL, UK
2
Institute for Microelectronics Technology, 142432 Chernogolovka, Russia
3
Manchester Centre for Mesoscience and Nanotechnology, University of Manchester, Oxford Road,
Manchester, M13 9PL, UK
4
TSC, The Ropewalk, Schofield Street, Heywood, Lancashire, OL10 1DS, UK
5
Graphene Industries Limited, The Fairbairn Building, PO Box 88, Manchester, M60 1QD, UK
6
High Field Magnet Laboratory, Radboud University Nijmegen, 6525 ED Nijmegen, Netherlands
7
School of Computer Science, University of Manchester, Oxford Road, Manchester, M13 9PL, UK
Received 30 July 2007, revised 22 August 2007, accepted 6 September 2007
Published online 8 November 2007
PACS 03.65.Pm, 73.20. r, 73.22. f, 73.43. f, 81.05.Uw
Graphene is the first example of truly two-dimensional crystals  it s just one layer of carbon atoms. It
turns out that graphene is a gapless semiconductor with unique electronic properties resulting from the
fact that charge carriers in graphene obey linear dispersion relation, thus mimicking massless relativistic
particles. This results in the observation of a number of very peculiar electronic properties  from an
anomalous quantum Hall effect to the absence of localization. It also provides a bridge between con-
densed matter physics and quantum electrodynamics and opens new perspectives for carbon-based elec-
tronics.
© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
1 Historical overview
Carbon demonstrates unusually complicated behavior, forming a number of very different structures. As
well as diamond and graphite, which have been known since ancient times, recently discovered fullere-
nes [1 3] and nanotubes [4] are currently a focus of attention for many physicists and chemists. Thus,
only 3-dimensional (diamond, graphite), 1-dimensional (nanotubes), and 0-dimensional (fullerenes)
allotropes of carbon were known. The two-dimensional form was conspicuously missing, resisting any
attempt at experimental observation  until recently.
The elusive two-dimensional form of carbon is named graphene, and, ironically, it is probably the
best-studied carbon allotrope theoretically. Graphene  planar, hexagonal arrangements of carbon atoms
 is the starting point for all calculations on graphite, carbon nanotubes, and fullerenes. At the same time,
numerous attempts to synthesize these two-dimensional atomic crystals have usually failed, producing
instead nanometer-size crystallites [5]. These difficulties are not surprising in light of the common belief
that truly two-dimensional crystals cannot exist [6 10] (in contrast to the numerous known quasi-two-
dimensional systems). Moreover, during synthesis, any graphene nucleation sites will have very large
perimeter-to-surface ratios, thus promoting collapse into other carbon allotropes.
*
Corresponding author: e-mail: kostya@manchester.ac.uk, Phone: +44 161 275 41 19, Fax: +44 161 275 40 56
© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
Original
Paper
phys. stat. sol. (b) 244, No. 11 (2007) 4107
In 2004, a very different and, at first glance, even naive approach to obtain graphene lead to a revolu-
tion in the field [11 13]. A single sheet (a monolayer of atoms) was extracted from 3-dimensional
graphite using a technique called micromechanical cleavage. Graphite is a layered material and can be
viewed as a number of two-dimensional graphene crystals weakly coupled together. By using this top-
down approach and starting with large, three-dimensional crystals, the researchers avoided all the issues
with the stability of small crystallites. Furthermore, the same technique has been used to obtain two-
dimensional crystals of other materials [11], such as boron-nitride, some dichalcogenides and high-
temperature superconductor Bi Sr Ca Cu O. This astonishing finding sends an important message:
two-dimensional crystals do exist and they are stable at ambient conditions.
Amazingly, this humble approach allows easy production of large (up to 100 µm in size), high-quality
graphene crystallites, and immediately triggered enormous experimental activity [14, 15]. Moreover, the
quality of the samples produced are so good that ballistic transport [12] and a quantum Hall effect (QHE)
[14, 15] can be observed easily. The former makes this new material a promising candidate for future
electronic applications, such as ballistic field-effect transistors (FETs). However, while this approach
suits all research needs [16], other techniques that provide a high yield of graphene are required for in-
dustrial production. Among the promising candidate, one should mention exfoliation of intercalated
graphitic compounds [17 21] and Si sublimation from SiC substrates [22, 23].
2 Linear spectra
From the point of view of its electronic properties, graphene is a two-dimensional zero-gap semiconductor
and its low-energy quasiparticles formally described by the Dirac-like Hamiltonian H =-i vFÃ  , where
vF ª 106m/s is the Fermi velocity, and à = (à , à ) are the Pauli matrices [24 27]. Neglecting many-body
x y
effects, this description is accurate theoretically and has also been proven experimentally by measuring the
energy-dependent cyclotron mass in graphene (which yields its linear energy spectrum) and, most clearly,
by the observation of a relativistic analogue of the integer QHE which will be discussed below.
The fact that charge carriers in graphene are described by the Dirac-like equation rather than the usual
Schrödinger equation can be seen as a consequence of graphene s crystal structure, which consists of two
equivalent carbon sublattices A and B. Quantum mechanical hopping between the sublattices leads to the
formation of two energy bands, and their intersection near the edges of the Brillouin zone yields the
conical energy spectrum near the  Dirac points K and K2 . As a result, quasiparticles in graphene exhibit
the linear dispersion relation E = kvF as if they were massless relativistic particles, with the role of the
speed of light played by the Fermi velocity vF H" c/300. Due to the linear spectrum, one can expect that
graphene s quasiparticles behave differently from those in conventional metals and semiconductors
where the energy spectrum can be approximated by a parabolic (free-electron-like) dispersion relation.
Although the linear spectrum is important, it is not the only essential feature that underpins the de-
scription of quantum transport in graphene by the Dirac equation. Above zero energy, the current carry-
ing states in graphene are, as usual, electron-like and negatively charged. At negative energies, if the
valence band is not completely filled, its unoccupied electronic states behave as positively charged qua-
siparticles (holes), which are often viewed as a condensed-matter equivalent of positrons. Note however
that electrons and holes in condensed matter physics are normally described by separate Schrödinger
equations, which are not in any way connected (as a consequence of the Seitz sum rule, the equations
should also involve different effective masses). In contrast, electron and hole states in graphene are inter-
connected, exhibiting properties analogous to the charge-conjugation symmetry in QED. For the case of
graphene, the latter symmetry is a consequence of its crystal symmetry because graphene s quasiparticles
have to be described by two-component wavefunctions, which is needed to define relative contributions
of sublattices A and B in the quasiparticles make-up. The two-component description for graphene is
very similar to the one by spinor wavefunctions in QED but the  spin index for graphene indicates
sublattices rather than the real spin of electrons and is usually referred to as pseudospin Ã.
There are further analogies with QED. The conical spectrum of graphene is the result of intersection
of the energy bands originating from sublattices A and B and, accordingly, an electron with energy E
www.pss-b.com © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
4108 K. S. Novoselov et al.: Electronic properties of graphene
propagating in the positive direction originates from the same branch of the electronic spectrum as the
hole with energy  E propagating in the opposite direction. This yields that electrons and holes belonging
to the same branch have pseudospin à pointing in the same direction, which is parallel to the momentum
for electrons and antiparallel for holes. This allows one to introduce chirality [27], that is formally a
projection of pseudospin on the direction of motion, which is positive and negative for electrons and
holes, respectively.
An alternative view on the origin of the chirality in graphene is based on the concept of  Berry phase .
Since the electron wave function is a two-component spinor, it has to change sign when the electron
moves along the closed contour. Thus the wave function gains an additional phase Ä„.
3 Electric field effect
In the absence of quality graphene wafers, most experimental groups are currently using samples ob-
tained by the original technique of micromechanical cleavage [11, 12], which provides high-quality crys-
tallites up to 1000 µm2 in size, which are sufficient for most research purposes and to prove concepts.
Superficially, the technique resembles nothing more sophisticated than drawing with a pencil [12] or
repeated peeling of graphite by adhesive tape [11] until the thinnest flakes are found. The critical ingre-
dient that made those efforts successful was that graphene becomes visible in an optical microscope if
placed on top of a Si wafer with a carefully chosen thickness of SiO2, owing to a very feeble interfer-
ence-like contrast with respect to an empty wafer [28]. If not for this simple yet effective way to scan
substrates in search of individual graphene crystallites, they would most probably remain undiscovered
today. Even if one would search for graphene ignoring its theoretical impossibility, more sophisticated
techniques usually employed for investigating atomically thin and small objects (such as scanning-probe
microscopy) have too low throughput to allow for finding of rare and isolated graphene crystals, whereas
scanning electron microscopy is also unsuitable because of the absence of clear signatures for the num-
ber of atomic layers. Note however that graphene was recently found to have a clear signature in Raman
microscopy [29, 30], which is likely to make this technique widely used for thickness inspection, al-
though potential crystals would still have to be identified optically first.
SiO2 also serves as an insulating layer, so a back-gate voltage can be applied to vary carrier concentra-
tion. Graphene is found to exhibit a pronounced ambipolar electric field effect (Fig. 1) such that charge
carriers can be tuned continuously between electrons to holes in concentrations n as high as 1013 cm 2
and their mobilities can exceed 10000 cm2/Vs and is practically temperature independent [12]. This
translates into ballistic transport on submicron scales. The room-temperature mobility is limited by im-
purities or corrugations of the graphene surface, which means that it can still be improved significantly,
perhaps up to 100.00 cm2/Vs.
Fig. 1 (online colour at: www.pss-b.com) Ambipolar electric field
effect in single-layer graphene. Positive (negative) Vg induce elec-
trons (holes) in concentrations n = Ä…Vg where the coefficient
Ä… H" 72 Ä„ 1010 cm 2/V for our field-effect devices with a 300 nm SiO2
.
layer used as a dielectric. The rapid decrease in resistivity with
adding charge carriers indicates their high mobility (in this case,
µ H" 5000 cm2/Vs and does not practically change with temperature
up to 300 K).
© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.pss-b.com
Original
Paper
phys. stat. sol. (b) 244, No. 11 (2007) 4109
Another important observation is that graphene s zero-field conductivity à does not disappear in the
limit of vanishing n but instead exhibits values close to the conductivity quantum e2/h per carrier type
[14]. We emphasize that it is the resistivity (conductivity) rather than the resistance (conductance) that is
quantized, in contrast to all other known quantum transport phenomena. Minimum quantum conductivity
has been predicted for Dirac fermions by a number of theories [31 37]. Some of them rely on a vanish-
ing density of states for the linear 2D spectrum. However, comparison between the behaviour of mass-
less and massive Dirac fermions in graphene and its bilayer allows one to distinguish between chirality-
and masslessness-related effects. To this end, bilayer graphene also exhibits a minimum conductivity of
the order of e2/h per carrier type which indicates that it is chirality, rather than the linear spectrum, that is
more important [38]. Moreover, most theories suggest Ámax = Ä„h/4e2, which is Ä„ times larger than the
typical value observed experimentally. This disagreement has become known as  the mystery of a miss-
ing pie , and it remains unclear whether it is due to approximations used by theory or because the ex-
periment probed only a limited range of possible sample parameters. In particular, experiments indicate
that, at low concentrations (n < 1011 cm-2), graphene conducts as a random network of electron and hole
puddles. Such microscopic inhomogeneity is inherent to the neutrality point but has so far not been taken
into account by theory. Furthermore, macroscopic inhomogeneity (on the scale larger than the mean free
path l) may also play an important role. Experiments show that, improving sample homogeneity by
thermal annealing or using smaller samples, changes Ámax so that it gets closer to h/4e2.
4 Quantum Hall effect
At this early stage, main experimental efforts have been focused on electronic properties in graphene,
trying to understand the consequences of its QED-like spectrum. Among the most spectacular phenom-
ena found in graphene so far are two new ( chiral ) quantum Hall effects.
Fig. 2 Hallmark of the QHE for massless Dirac fermions is plateaux
in à (top panel) at half integers of 4e2/h whereas Áxx (bottom panel)
xy
vanishes for the same carrier concentrations n. T = 4 K; B = 14 T.
www.pss-b.com © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
4110 K. S. Novoselov et al.: Electronic properties of graphene
Fig. 3 Room-temperature quantum Hall effect. Because
quasiparticles in graphene behave as massless Dirac fer-
mions and exhibit little scattering even under ambient
conditions, the QHE survives up to room T. The dotted
curve is the Hall conductivity à that exhibits clear pla-
xy
teaux at 2e2/h for both electrons and holes. The longitudi-
nal conductivity Áxx (solid curve) reaches zero at the same
gate voltages.
Figure 2 shows the QHE behaviour observed in graphene. QHE in single layer graphene shows up as an
uninterrupted ladder of equidistant steps in Hall conductivity à which persists through the neutrality
xy
(Dirac) point, where charge carriers change from electrons to holes (Fig. 2). The sequence is shifted with
respect to the  standard QHE sequence by 12, so that à =ą4e2/h(N + 1/2) where N is the Landau
/
xy
level (LL) index and factor 4 appears due to double valley and double spin degeneracy. The QHE has
been dubbed  half-integer to reflect both the shift and the fact that, although it is not a fractional QHE,
it is not the integer QHE either. The unusual sequence is now well understood as arising due to the
QED-like quantization of graphene s electronic spectrum in magnetic field B, which is described by
EN =Ä…vF 2e BN where sign Ä… refers to electrons and holes [26, 27, 39 41]. The existence of a quantized
level at zero E, which is shared by electrons and holes, is essentially everything one needs to know to
explain the anomalous QHE sequence [38, 40, 41]. An alternative explanation for the half-integer QHE
is to invoke the coupling between pseudospin and orbital motion, which gives rise to a Berry phase of Ä„
accumulated along cyclotron trajectories [14, 15]. The additional phase leads to a half-period shift in the
phase of quantum oscillations and, in the QHE limit, to a half-step shift [14, 15].
Graphene linear spectra and the large value of the Fermi velocity ensures huge orbital splitting. The
energy gap between N = 0 and N =Ä…1 Landau levels is given by "E H" 400 (K) B , where B is magnetic
field in Tesla. This implies that at B = 30 T the splitting is of the order of 2200 K, almost an order of
magnitude higher than the room temperature. Moreover, temperature independent high mobility ensures
that the high field limit µ Ę% B 1 is satisfied in a modest field of a few Tesla. These led to observation of
QHE at room temperatures [42] (Fig. 3), which is very promising for metrology applications.
5 Conclusions
Graphene is the first example of a truly two-dimensional crystal. This opens many interesting questions
concerning the thermodynamics, lattice dynamics, and structural properties of such systems. Being a
gapless semiconductor with a linear energy spectrum, single-layer graphene realizes a two-dimensional,
massless Dirac fermion system that is of crucial importance for understanding unusual electronic proper-
ties, such as an anomalous QHE, absence of the Anderson localization, etc. These peculiarities are im-
portant for developing new electronic devices.
Acknowledgements K.S.N. is grateful to the Royal Society for financial support. S.V.M. acknowledges support
from RFBR05-02-17638,  Quantum MacroPhysics and  Quantum Nanosctructures programs. D. C. E. would like
to thank CNPq-Brazil. Measurements in high magnetic field were supported by the Foundation for Fundamental
Research on Matter and the EU (EuroMagNet EU RII3-CT-2004-506239).
© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.pss-b.com
Original
Paper
phys. stat. sol. (b) 244, No. 11 (2007) 4111
References
[1] R. F. Curl, Rev. Mod. Phys. 69, 691 (1997).
[2] H. Kroto, Rev. Mod. Phys. 69, 703 (1997).
[3] R. E. Smalley, Rev. Mod. Phys. 69, 723 (1997).
[4] S. Iijima, Nature 354, 56 (1991).
[5] C. Oshima and A. Nagashima, J. Phys.: Condens. Matter 9, 1 (1997).
[6] R. E. Peierls, Helv. Phys. Acta 7, 81 (1934).
[7] R. E. Peierls, Ann. Inst. Henri Poincare 5, 177 (1935).
[8] L. D. Landau, Phys. Z. Sowjet Union 11, 26 (1937).
[9] L. D. Landau and E. M. Lifshitz, Statistical Physics, Part I (Pergamon Press, Oxford, 1980).
[10] N. D. Mermin, Phys. Rev. 176, 250 (1968).
[11] K. S. Novoselov et al., Proc. Natl. Acad. Sci. 102, 10451 (2005).
[12] K. S. Novoselov et al., Science 306, 666 (2004).
[13] A. K. Geim and K. S. Novoselov, Nature Mater. 6, 183 191 (2007).
[14] K. S. Novoselov et al., Nature 438, 197 (2005).
[15] Y. Zhang et al., Nature 438, 201 (2005).
[16] www.grapheneindustries.com
[17] M. S. Dresselhaus and G. Dresselhaus, Adv. Phys. 51, 1 (2002).
[18] H. Shioyama, J. Mater. Sci. Lett. 20, 499 (2001).
[19] L. M. Viculis et al., Science 299, 1361 (2003).
[20] S. Horiuchi et al., Appl. Phys. Lett. 84, 2403 (2004).
[21] S. Stankovich et al., J. Mater. Chem. 16, 155 (2006).
[22] A. J. Van Bommel, J. E. Crombeen, and A. Van Tooren, Surf. Sci. 48, 463 (1975).
[23] C. Berger et al., J. Phys. Chem. B 108, 19912 (2004).
[24] P. R. Wallace, Phys. Rev. 71, 622 (1947).
[25] J. C. Slonczewski and P. R. Weiss, Phys. Rev. 109, 272 (1958).
[26] J. W. McClure, Phys. Rev. 104, 666 (1956).
[27] F. D. M. Haldane, Phys. Rev. Lett. 61, 2015 (1988).
[28] P. Blake, K. S. Novoselov, A. H. Castro Neto, D. Jiang, R. Yang, T. J. Booth, A. K. Geim, E. W. Hill,
arXiv:0705.0259, to be published in Appl. Phys. Lett. (2007).
[29] A. C. Ferrari, J. C. Meyer, V. Scardaci, C. Casiraghi, M. Lazzeri, F. Mauri, S. Piscanec, D. Jiang, K. S. Novose-
lov, S. Roth, and A. K. Geim, Phys. Rev. Lett. 97, 187401 (2006).
[30] D. Graf, F. Molitor, K. Ensslin, C. Stampfer, A. Jungen, C. Hierold, and L. Wirtz, Nano Lett. 7, 238 (2007).
[31] M. I. Katsnelson, Eur. Phys. J. B 51, 157 (2006).
[32] J. Tworzydlo, B. Trauzettel, M. Titov, A. Rycerz, and C. W. J. Beenakker, Phys. Rev. Lett. 96, 246802 (2006).
[33] E. Fradkin, Phys. Rev. B 33, 3263 (1986).
[34] P. A. Lee, Phys. Rev. Lett. 71, 1887 (1993).
[35] A. W. W. Ludwig, M. P. A. Fisher, R. Shankar, and G. Grinstein, Phys. Rev. B 50, 7526 (1994).
[36] K. Ziegler, Phys. Rev. Lett. 80, 3113 (1998).
[37] K. Nomura and A. H. MacDonald, Phys. Rev. Lett. 96, 256602 (2006).
[38] K. S. Novoselov, E. McCann, S. V. Morozov et al., Nature Physics 2, 177 (2006).
[39] Y. Zheng and T. Ando, Phys. Rev. B 65, 245420 (2002).
[40] V. P. Gusynin and S. G. Sharapov, Phys. Rev. Lett. 95, 146801 (2005).
[41] N. M. R. Peres, F. Guinea, and A. H. Castro Neto, Phys. Rev. B 73, 125411 (2006).
[42] K. S. Novoselov, Z. Jiang, Y. Zhang, S. V. Morozov, H. L. Stormer, U. Zeitler, J. C. Maan, G. S. Boebinger,
P. Kim, and A. K. Geim, Science 315, 1379 (2007).
www.pss-b.com © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


Wyszukiwarka